2012 ready, set, fuse! the coronavirus spike protein and acquisition of fusion competence

24
Viruses 2012, 4, 557-580; doi:10.3390/v4040557 viruses ISSN 1999-4915 www.mdpi.com/journal/viruses Review Ready, Set, Fuse! The Coronavirus Spike Protein and Acquisition of Fusion Competence Taylor Heald-Sargent and Tom Gallagher * Department of Microbiology and Immunology, Loyola University Medical Center, 2160 South First Avenue, Maywood, IL 60153, USA; E-Mail: [email protected] * Author to whom correspondence should be addressed; E-Mail: [email protected]; Tel.: +1-708-216-4850. Received: 15 March 2012; in revised form: 29 March 2012 / Accepted: 2 April 2012 / Published: 12 April 2012 Abstract: Coronavirus-cell entry programs involve virus-cell membrane fusions mediated by viral spike (S) proteins. Coronavirus S proteins acquire membrane fusion competence by receptor interactions, proteolysis, and acidification in endosomes. This review describes our current understanding of the S proteins, their interactions with and their responses to these entry triggers. We focus on receptors and proteases in prompting entry and highlight the type II transmembrane serine proteases (TTSPs) known to activate several virus fusion proteins. These and other proteases are essential cofactors permitting coronavirus infection, conceivably being in proximity to cell-surface receptors and thus poised to split entering spike proteins into the fragments that refold to mediate membrane fusion. The review concludes by noting how understanding of coronavirus entry informs antiviral therapies. Keywords: coronavirus; virus entry; viral pathogenesis; spike protein; carcinoembryonic antigen; angiotensin converting enzyme 2; endocytosis; cathepsin; transmembrane protease; membrane fusion 1. Introduction Coronaviruses (CoVs) are enveloped RNA viruses causing respiratory and enteric diseases. The CoVs are widely distributed in nature and their zoonotic transmissions into human populations can cause epidemic disease. After entering into respiratory or gastrointestinal tracts, these viruses establish OPEN ACCESS

Upload: others

Post on 11-Sep-2021

1 views

Category:

Documents


0 download

TRANSCRIPT

Page 1: 2012 Ready, Set, Fuse! The Coronavirus Spike Protein and Acquisition of Fusion Competence

Viruses 2012, 4, 557-580; doi:10.3390/v4040557

virusesISSN 1999-4915

www.mdpi.com/journal/viruses

Review

Ready, Set, Fuse! The Coronavirus Spike Protein and Acquisition of Fusion Competence

Taylor Heald-Sargent and Tom Gallagher *

Department of Microbiology and Immunology, Loyola University Medical Center,

2160 South First Avenue, Maywood, IL 60153, USA; E-Mail: [email protected]

* Author to whom correspondence should be addressed; E-Mail: [email protected];

Tel.: +1-708-216-4850.

Received: 15 March 2012; in revised form: 29 March 2012 / Accepted: 2 April 2012 /

Published: 12 April 2012

Abstract: Coronavirus-cell entry programs involve virus-cell membrane fusions mediated

by viral spike (S) proteins. Coronavirus S proteins acquire membrane fusion competence

by receptor interactions, proteolysis, and acidification in endosomes. This review describes

our current understanding of the S proteins, their interactions with and their responses to

these entry triggers. We focus on receptors and proteases in prompting entry and highlight

the type II transmembrane serine proteases (TTSPs) known to activate several virus fusion

proteins. These and other proteases are essential cofactors permitting coronavirus infection,

conceivably being in proximity to cell-surface receptors and thus poised to split entering

spike proteins into the fragments that refold to mediate membrane fusion. The review

concludes by noting how understanding of coronavirus entry informs antiviral therapies.

Keywords: coronavirus; virus entry; viral pathogenesis; spike protein; carcinoembryonic

antigen; angiotensin converting enzyme 2; endocytosis; cathepsin; transmembrane protease;

membrane fusion

1. Introduction

Coronaviruses (CoVs) are enveloped RNA viruses causing respiratory and enteric diseases. The

CoVs are widely distributed in nature and their zoonotic transmissions into human populations can

cause epidemic disease. After entering into respiratory or gastrointestinal tracts, these viruses establish

OPEN ACCESS

Page 2: 2012 Ready, Set, Fuse! The Coronavirus Spike Protein and Acquisition of Fusion Competence

Viruses 2012, 4 558

themselves by entering and infecting lumenal macrophages and epithelial cells. The cell entry

programs for these viruses are orchestrated by the viral spike (S) proteins that bind cellular receptors

and also mediate virus-cell membrane fusions. CoV diversity is reflected in the variable S proteins,

which have evolved into forms differing in their receptor interactions and their response to various

environmental triggers of virus-cell membrane fusion. As such, seemingly minor differences in CoV S

protein structure and function often correlate with striking changes in CoV tropism and virulence. This

review focuses on the S: receptor interactions and events leading to membrane fusion and successful

cell entry.

The S proteins are large; an average CoV S protein contains 1,300 amino acids and 20 asparagine-linked

glycans, making the S proteins substantially larger than the glycoproteins mediating receptor binding

and membrane fusion of other enveloped viruses. These S monomers are assembled into trimers within

the ER of virus-producing cells [1] and incorporated into virus particles at the CoV budding sites, at or

near the ER-Golgi Intermediate Compartment [2]. For the human Severe Acute Respiratory Syndrome

(SARS) CoV and the Mouse Hepatitis Virus (MHV) CoV, the virion-associated S proteins have been

imaged at ~20 angstrom resolution by cryo-electron microscopy [3]. There are about 80 S trimers on

each SARS-CoV or MHV particle [4,5]. Each S trimer is held onto viruses by transmembrane anchors

located near C-termini, with the vast majority of the S protein mass in extra-virion locations,

protruding about 20 nm from the limiting virus membrane of each 80 nm-diameter particle [4]

(Figure 1A).

Atomic–resolution structures are known for domains comprising ~20% of the SARS and MHV S

proteins. These include the receptor binding domains (RBDs), which are part of the N-terminal “S1”

half of the primary S sequence (Figure 1B). RBD structures in complex with cognate cell receptors

(Figure 1C and 1D) have provided several insights into the initial virus-cell interaction [6-9]. The other

atomically-resolved structures amount to a small portion of the C-terminal half of the primary S

sequence (Figure 1E) and reveal a coiled-coil made of three alpha helices termed heptad repeat (HR) 1

along with three chains termed HR2 that are bound onto the outer surfaces of the coiled coil (Figure

1E; note the single HR1-HR2 heteromer for clarification). This complex of three HR1s and three HR2s

is termed the six-helix bundle (6-HB) and is a very stable, protease-resistant structure that is part of S

proteins after they have denatured, or after they have catalyzed membrane fusion

(“post-fusion” form) [10-13]. The structures of HR1 and HR2 in the native or “pre-fusion” S

conformations are not known. Furthermore, atomic structures of the remaining ~80% of S proteins are

not known for any of the S protein conformations.

Figure 1. CoV virion and spike protein features. (A) Depiction of the virion. The virion

core includes a helical ribonucleoprotein core consisting of a ~30 kilobase single-stranded

RNA that is surrounded by nucleocapsid (N) proteins. The virion membrane is enriched

with viral membrane proteins and includes a small number of envelope proteins. The

spikes protrude about 20 nm from the limiting virion membrane; (B) Depiction of the S

protein. A single S protein is depicted as a rectangle, from N- to C-terminus in left-to-right

orientation. Relevant structural features are higlighted as follows: N-terminal receptor

binding domain (N-RBD) in dark blue, with receptor binding motif (RBM) in yellow;

C-RBD in brown, with RBM in yellow; cleavage sites (CS) 1 and 2, fusion peptide (FP) in

Page 3: 2012 Ready, Set, Fuse! The Coronavirus Spike Protein and Acquisition of Fusion Competence

Viruses 2012, 4 559

red, heptad repeat (HR) regions 1 and 2 in green, with N and C-termini in yellow;

transmembrane span ( TM ) depicted as membrane bilayer; cytoplasmic tail (CT) in light

blue. The image is adjusted to the primary amino acid scale, which is 1255 residues for the

SARS-CoV S protein. Note that the RBDs that are depicted come from different CoV S

proteins; N-RBD from MHV and C-RBD from SARS; (C) Structure of the MHV N-RBD in

complex with its CEACAM receptor (PDB 3R4D; reference [8]). The alpha carbon

structure of the MHV N-RBD (in blue) is highlighted by RBMs (in yellow). The RBMs

contact the N-terminal CEACAM receptor immunoglobulin domain (in purple); (D) Structure

of the SARS C-RBD in complex with its ACE2 receptor (PDB 2AJF; reference [7]). The

SARS C-RBD (in brown) is highlited by RBMs (in yellow). The RBMs contact a

virus-binding hotspot [14] on the ACE2 receptor (in purple); (E) Structure of the

postfusion HR1-HR2 bundle (PBD 1WYY; reference [15]). The HR1-HR2 bundle is

depicted as alpha carbon tracings. Each of the three HR1-HR2 are a distinguished by color.

To enhance the display, a single HR1-HR2 is extracted from the image and included below

the 6-stranded bundle. This HR1-HR2 includes highlights (in yellow) depicting the N-

terminal end of the HR1 and the C-terminal end of the HR2. The postfusion configuration

is shown as it might appear relative to the fused membrane, when the fusion peptide (not

shown) would extend into the fused membrane from the N-terminal end of HR1 and the

HR2 would be held into the fused membrane by the TM span.

Page 4: 2012 Ready, Set, Fuse! The Coronavirus Spike Protein and Acquisition of Fusion Competence

Viruses 2012, 4 560

Models depicting the S-mediated membrane fusion event have extended from knowledge of S

protein structures and functions. In part, these models are deemed reasonable because the postfusion 6-

HB conformations in SARS and MHV S proteins are so strikingly similar to postfusion forms of

influenza HA2, paramyxovirus F2, Ebolavirus GP2 and HIV gp41 [16]. In analogy to these more

widely-studied and well-understood viral fusion proteins, the CoV S-mediated membrane fusion

process is generally viewed as schematized in Figure 2. Following receptor binding, the CoV S

proteins undergo conformational change that exposes hydrophobic domains that likely include a

so-called fusion peptide (FP), which embeds into the target cell membrane. Primary sequence data,

along with measured fusion activities of site-directed S mutants, suggest that the FPs are near or

immediately N-terminal to the HR1 ([17,18]; see Figure 1 for proposed FP location). At this stage in

which the FP is in the target cell membrane and the transmembrane (TM) span is in the virion

membrane, the S proteins are considered to be in “fusion intermediate” conformations. What follows is

the refolding of the regions between the FPs and the TM spans into the postfusion 6-HB configuration,

a likely multi-step process that brings opposing membranes into sufficient proximity for lipid mixing

and ultimate coalescence. In addition to FP, HR1, HR2 and TM spans, structural elements participating

in these later steps include a hydrophobic tryptophan-rich region at the boundary between TM spans

and ectodomains [19,20] and a post-translationally palmitoylated region at the boundary between TM

spans and cytoplasmic domains [21-23]. At the completion of membrane fusion, FPs and TMs are

hypothesized to be adjacent to each other, extending from the same end of the 6-HBs and into the

membrane bilayer [13,24,25]. In clear analogy to the majority of viruses promoting membrane fusion

in this way “HR2” peptides that bind to the HR1 regions of fusion intermediate conformations will

potently block CoV cell entry at the membrane fusion stage, most likely by preventing the completion

of late-stage refolding into 6-HBs [26-28].

This understanding of CoV entry has provided a sophisticated basis for continued exploration of the

CoVs. Incentives to continue research on CoV entry are clear–with proven CoV zoonotic potential and

extraordinary pathogenicity in humans, a knowledge base of CoV entry offers potential to save lives.

There are significant gaps in this knowledge base, and current incentives are to further understand

virus-receptor interactions and S protein triggering to catalyze membrane fusion. In adding to the

knowledge base, the CoV model system will also bring out new insights of general relevance to virus

entry and the membrane fusion. This review covers recent progress in adding to the fund of knowledge

on CoV entry.

The sections of this review are organized according to the events of the CoV infection cycle,

beginning with the biogenesis and secretion of virus particles. This starting point is chosen to highlight

the S protein maturation events in virus-producing cells and to emphasize the continuous preparation

for membrane fusion that takes place both in virus-producing and virus-target cells. The subsequent

sections proceed to virus-receptor interactions and their roles in inducing S protein conformational

changes. Next, the importance of endocytosis is discussed, and S proteolytic scissions at the cell

surface and within endosomes are then underscored as critical features of CoV cell entry.

Page 5: 2012 Ready, Set, Fuse! The Coronavirus Spike Protein and Acquisition of Fusion Competence

Viruses 2012, 4 561

Figure 2. Schematic illustration of CoV S protein-mediated membrane fusion. The

illustrations represent several steps of S protein conformational changes that may take

place during membrane fusion. In the first step, receptor binding, pH reduction and/or S

protein proteolysis induces dissociation of S1 from S2. This step is documented for some

MHVs [29,30]. In the second step, the fusion peptide (FP) is intercalated into the host cell

membrane. This is the fusion-intermediate stage. In the third stage, the part of the S protein

nearest to the virus membrane refolds onto a heptad repeat 1 (HR1) core to form the

six-helix bundle (6-HB), which is the final postfusion configuration of the S2 protein.

2. S protein Morphogenesis and Initial S Protein Proteolysis

The events that ready viruses for cell entry begin in the virus-producing cell. For the CoVs, this

beginning stage is in the endoplasmic reticulum of virus-producing cells, where S proteins are

synthesized and assembled into trimers. This ER-localized assembly into trimers is a slow process,

taking approximately 15 min to generate the greater than 500-kilodalton glycoprotein complexes [1].

Nascent S trimers then transit through the exocytic pathway, embedded by their C-terminal

transmembrane domains to vesicle membranes, or to virion membranes if incorporated into viruses at

the virus budding sites [2], ultimately residing on the plasma membrane or on secreted virions. It is

conceivable that the CoV infection process establishes novel exocytic routes, perhaps with distended

organelles able to accommodate secretion of very large S and CoV virion cargo [31]. Exocytic vesicles

may even be altered so that the virions and S proteins they house are preserved and properly routed

out of the cell. Of note here, the CoV E proteins, small oligomeric transmembrane peptides

that accumulate on ER-Golgi intermediate compartment membranes [32] and possess ion channel

activities [33], may have a general role in corrupting secretory organelles so that virions are expelled in

infectious forms [34].

Page 6: 2012 Ready, Set, Fuse! The Coronavirus Spike Protein and Acquisition of Fusion Competence

Viruses 2012, 4 562

The virus-producing cell is where the S proteins of some CoVs undergo an early proteolysis (event

“1” in Figure 3). The trans-Golgi network houses acid pH-activated serine proteases, notably furin [35],

that cleave some S proteins at positions immediately C-terminal to multibasic cleavage sites, at a place

that we labeled as cleavage site 1 (CS-1) in Figure 1. Scission at CS-1, also designated as S1-S2

cleavage, separates the RBDs and fusion machineries into noncovalently-associated peripheral S1and

transmembrane-anchored S2 fragments. That furin proprotein convertases are the responsible proteases

has been established by using peptide furin inhibitors, which arrest the S cleavages [36]. Notably, only

a subset of CoVs have S proteins with the multibasic cleavage sites, and of these, only a select few

have the consensus K/R-X-K/R-K/R motif at the P4-P1 position. Thus, in most CoV infections, S

protein populations are left intact or are only partially cleaved in virus-producing cells. For those S

proteins that do harbor the cleavage site motifs, site-directed mutation of the multibasic residues to

render “uncleavable” states does not destroy virus infectivity [37,38]. Experiments involving

pharmacologic inhibition of furin complement these evaluations of mutant viruses, and reveal that

reduced furin activities greatly reduce S cleavages without diminishing virus infectivities [36].

Figure 3. Proteolytic events during the CoV infection cycle. S cleavage events may take

place during virus egress from producer cells (1), extracellular transit to target cells (2), on

the target cell surface (3), and within the acidified endosome (4).

S1-S2 cleavage does, however, increase the potency of CoV S proteins to mediate cell-cell

fusions [39,40]. This influence on syncytial developments may explain some of the evolutionary

diversity of S1-S2 cleavages. Syncytia may benefit CoVs in certain niches, i.e., during acute

infections, by promoting rapid virus dissemination, and this proviral effect may select for the

multibasic cleavage motifs. In other niches, i.e., during chronic or smoldering infections, weakened

Producer Cell

Target Cell

1

2

3

4

Page 7: 2012 Ready, Set, Fuse! The Coronavirus Spike Protein and Acquisition of Fusion Competence

Viruses 2012, 4 563

syncytial cells may have greatly compromised virus-producing capacities, and this antiviral effect may

select against the same multibasic peptide sequences. Additional, entirely distinct selective forces may

target the basic peptide motifs, as the K/R-X-K/R-K/R stretch can mediate virus binding to heparan

sulfate, a known receptor for murine CoVs [41]. Again, adsorption to heparan sulfate can be proviral in

tissue culture and antiviral in vivo, making for complex selective forces impinging on multibasic S

peptide motifs.

Exactly how S1-S2 cleavages promote syncytia (cell-cell fusion), often without significant effects

on virus entry (virus-cell fusion), are not yet entirely clear. Notably, the S1-S2 cleavage site is distant

from any hydrophobic peptides, hundreds of residues from the presumed fusion peptide (Figure 1).

This positioning is clearly unlike the analogous cleavage sites on influenza hemagglutinin and HIV

gp160, which are adjacent to fusion peptides and thus create hydrophobic N-termini for target

membrane insertion. With simple models of S1-S2 cleavage liberating an N-terminal fusion peptide

appearing unsatisfactory, there have been hypotheses for additional S cleavages nearer to the fusion

peptides. Support for this hypothesis that S is a pre-pro-protein is now in place. As these cleavages to

the final fusion-active S likely take place at later times in the continuum of virus assembly, secretion

and entry into new cells, they are described in detail in the following sections.

3. Engaging the CoV Receptors

All CoVs, regardless of their pre-primed S protein cleavage status, initiate new infections by

binding receptors on virus-target cells. These receptor-binding events, their variabilities amongst the

CoVs, and their potential for priming subsequent virus-target cell membrane fusions, are discussed here.

A fascinating feature of the CoVs is in the diversity of receptor-binding events. This diversity

shows up readily upon inspection of the CoV receptors, which include integral-membrane proteins and

sugars that, as a group, have no obvious structural similarities (Table 1). This variation presents

interesting questions regarding which receptors are simply mediating CoV adsorption and which might

be doing more; i.e., generating fusion-promoting S protein conformational changes and/or directing

viruses through particular endocytic routes. Several studies have demonstrated that CoV receptors may

not require special endocytic functions beyond S binding: Integral-membrane protein receptors lacking

cytoplasmic tails and thus unlikely to regulate virus endocytosis or transduce virus-promoting signals

will still provide for virus entry [42,43]. In the case of MHV-CoV and its CEACAM receptors

(Table 1), cell entry can be catalyzed by soluble, exogenously-added CEACAM ectodomains [44-46],

indicating that this receptor can advance entry through S binding, and that a role for the receptor in

guiding bound viruses into particular endocytic organelles is not absolutely essential.

Diversity of receptor interaction is also readily evident by examining the CoV S proteins. The S

proteins are striking in that, as a group, they possess at least two RBDs. One RBD encompasses the S1

N-terminal domain (N-terminal RBD in Figure 1 and Table 1). Recent structural resolution of this

NTD from MHV-CoV has revealed a galectin-like beta-sandwich fold, and indeed, the NTDs from

related CoV strains (human CoV-OC43 and bovine CoV) are functional lectins and do bind various

acetylated sialic acids [8]. Remarkably, the resolved MHV NTD structure lacks a peptide loop

necessary for sugar binding, and thus does not operate as a lectin, instead possessing high affinity for

the CEACAM proteins. The intriguing indications here are that the N-terminal RBD structures can

Page 8: 2012 Ready, Set, Fuse! The Coronavirus Spike Protein and Acquisition of Fusion Competence

Viruses 2012, 4 564

evolve disparate sugar or protein receptor-binding activities. The findings are valuable in considering

CoV receptor switching as it relates to host range and zoonotic infection potential.

Table 1. Coronaviruses and their receptors. Viruses include TGEV: Transmissible

Gastroenteritis Virus, PRCoV: Porcine Respiratory Coronavirus, HCoV-229E: Human

Coronavirus 229E, HCoV-NL63: Human Coronavirus Netherlands-63, SARS-CoV: Severe

Acute Respiratory Syndrome-related Coronavirus, MHV: Murine Hepatitis Virus,

HCoV-OC43: Human Coronavirus OC-43, BCoV: Bovine Coronavirus. Cell receptors for

N-terminal virus RBDs include sugars and CEACAM1: Carcinoembryonic Antigen-related

Cell Adhesion Molecule 1. Cell receptors for C-terminal RBDs include APN:

Aminopeptidase-N, and ACE2: Angiotensin Converting Enzyme 2.

The other CoV RBD encompasses a region near the middle of S1 and has been termed the

C-terminal domain (CTD). The C-terminal RBD structures of human SARS (shown in Figure 1D) and

NL63 (not shown) CoVs are completely distinct, a beta sheet architecture in SARS and a beta

sandwich in NL63, yet these two CTD RBDs bind to the same region within the ACE2 receptor,

termed the “virus binding hotspot” [6,8,14,47]. In the SARS S and CTD, there are well-described

mutations that change affinities for human, civet cat and bat ACE2 receptors [9,47,48], and as

expected, virus affinity for the human ACE2 correlates with human infection and epidemic human

to human transmission [47]. The CTD structures on those CoVs that bind APN (or other as yet

undetermined receptors) are not yet available, but once determined, may foster analogous identification

Page 9: 2012 Ready, Set, Fuse! The Coronavirus Spike Protein and Acquisition of Fusion Competence

Viruses 2012, 4 565

of S mutations that change affinities for far more diverse receptors. Such studies will promote

understanding of how CoV C-terminal RBDs, like the N-terminal RBDs, possess flexible architectures

that can evolve disparate receptor-binding activities, again contributing to acquisition of new host

ranges and zoonoses.

Notably, the existence of two CoV RBDs can make one dispensable. The MHVs, having

CEACAM-binding sites on NTDs, readily delete CTD segments during passage in tissue culture [49,50].

These deletion-mutations have profound influences on CoV virulence, greatly decreasing the murine

virus neurovirulence [51]. Another interesting example is with the porcine CoVs. Transmissible

gastroenteritis virus (TGEV), with two functional RBDs (Table 1), can infect both respiratory and

enteric tracts [52]. However, naturally-occurring mutants with substitutions or deletions in N-terminal

RBDs lack lectin activities [53] and are strictly limited to the respiratory tract and therefore are

attenuated [54,55]. Thus a dual RBD architecture may correlate with in vivo tropism and virulence,

likely due to a complexity of multiple receptor interactions that the CoVs employ in the natural

animal environment.

Diverse receptor interactions precede the CoV S-mediated fusion reactions, and while the binding

of relatively low-affinity carbohydrate receptors may not generate fusion-promoting S protein

conformations protein receptors that bind S proteins at high affinity clearly do, as evidenced most

extensively by studies with MHV. Early seminal findings using MHV demonstrated that alkaline pH

increased S fusion activities and S1 release, a readily observed conformational change [29]. Soluble

CEACAM receptors were then found to catalyze S1 release [30,56], and biological relevance was

subsequently established when soluble receptors were found to support infectious MHV entry into

CEACAM-negative cells [44]. More recently, using an MHV2 strain, soluble CEACAM receptors

generated SDS-resistant S protein trimers with unique lipophilicities and protease susceptibilities [57].

Thus the MHV model system divulges relatively stable CEACAM receptor-induced S conformations

that are quite likely the intermediate structures on the way to membrane fusion (see Figure 2 for

hypothetical illustration of receptor-induced generation of fusion intermediate S structures).

What is not disclosed by the MHV model system, however, is how CEACAM binding to the NTD

RBDs can uncover the fusion machinery in S2. In the primary S sequence, the NTD RBDs are distant

from the fusion-inducing peptides. Structural biologists will undoubtedly address this issue most

effectively, but at present, intriguing molecular genetic data strongly suggest connections between

RBDs and fusion apparati in the context of the native S trimers. One of the first findings in support of

such connections was with the identification of a mutation in the fusion domain that destroyed an

antibody epitope in the RBD [58]. There have been numerous comparable observations since then.

Indeed, MHV evolution, both in vitro and in vivo, frequently fixes substitution mutations in or near the

RBDs, and into the fusion regions [59-61]. In several instances, these mutations do not change RBD

affinities for CEACAMs, but do increase or decrease the tendency for S proteins to undergo

CEACAM-induced transitions into fusion-intermediate forms [57,62], suggesting that one selective

pressure relates to MHV responsiveness to the CEACAM receptor “trigger”. Finally, the fact that these

mutations are frequently associated with reduced virus virulence creates the additional suggestion that

any complete understanding of CoV disease requires appreciation of how receptors operate as

fusion triggers.

Page 10: 2012 Ready, Set, Fuse! The Coronavirus Spike Protein and Acquisition of Fusion Competence

Viruses 2012, 4 566

Relative to CEACAM and its effect on MHVs, the roles of ACE2 and APN receptors in triggering

structural transitions of human CoVs is less clear. It is, for example, uncertain whether ACE2 binding

to SARS or NL63 S proteins dissociates RBDs from fusion modules and exposes hydrophobic “fusion

peptides”. Compelling data do, however, suggest that ACE2-bound S proteins are set up for proteolytic

scissions that activate membrane fusion, as described in the following sections. The human CoV

receptors, therefore, may be similar to the murine CEACAM CoV receptors in that their binding

energies drive S proteins into intermediate forms with exposed proteolytic substrate sites. Host

proteases may then ultimately liberate the fusion machineries and allow them to operate in CoV entry.

Finally, there is the question of whether virus entry and the receptor-induced conformational

changes leading to membrane fusion take place at the plasma membrane or at a later time, after viruses

have entered into endosomes. This question is explored in the following section.

4. Endocytosis Following Coronavirus: Cell Binding

Endocytosis is the principal cell entry route for most viruses [63]. Receptor-mediated endocytosis is

an advantageous entry route as it can allow the virus (in endosomes) to pass through a barrier of

cortical actin and traffic deeply into the cell before fusing and expelling its genome to cytosol [64].

The endosome also bathes the virus in an acidic, proteolytic environment that is frequently necessary

to activate viral fusion. It appears that the majority of CoVs enter cells via endocytosis. However, a

subset of CoVs have been thought to enter in a more straightforward direct fusion with the plasma

membrane, making the CoVs useful for dissecting various cell entry routes.

The idea that some CoVs do not endocytose but rather enter by fusion directly at the plasma

membrane comes from several findings, including the fact that several CoVs readily induce syncytia,

i.e., plasma membrane-localized cell-cell fusions. In those studies where pH effects were investigated,

syncytial developments increased in slightly alkaline media [56], suggesting no need for endosomal

acid exposures to activate fusions. Similarly, some CoVs, notably the JHM strain of MHV, are impervious

to inhibitors of endosomal acidification [65]. Finally, when incubated with receptor-bearing cells at

extraordinarily high input multiplicities, some CoVs can induce immediate cell-cell fusions, often

called “fusion-from-without” [66]. But none of these observations necessarily confirm that CoVs

naturally enter cells via fusion at the plasma membrane. First, it may be inappropriate to equate

syncytial formation (cell-cell fusion) with virus entry (virus-cell fusion), because these two processes

have distinct requirements, as revealed by S mutations that abolish syncytial formation without

affecting virus infectivity, as discussed later in this review. Second, increased syncytial activity at

basic pH may reflect the pH optima of the cell-surface trypsin-like proteases that create fusion-active S

fragments, and may primarily apply to fusions mediated by S proteins that are not associated with

virions. Third, the resistance of some CoVs to endosome neutralization can indicate fusion in early

non-acidic endosomes, but cannot be used to conclude that virus-cell fusion takes place at the plasma

membrane. Indeed there are several reports, particularly in HIV research, that viruses acquiring

membrane fusion competence at neutral pH may still utilize endosomal networks during entry [67].

This leaves the “fusion from without” phenomena, clearly plasma membrane-based virus-cell fusion

events, but perhaps occurring with very low frequencies and thus only observed in artificial

experimental conditions in which viruses are applied to cells at extremely high doses.

Page 11: 2012 Ready, Set, Fuse! The Coronavirus Spike Protein and Acquisition of Fusion Competence

Viruses 2012, 4 567

That the CoVs generally utilize cellular endocytic machinery during entry is also supported by

studies aimed at dissecting CoV entry pathways [68]. Many of these investigations assess the efficiencies

of CoV infections after specifically arresting clathrin, or caveolar, or cholesterol-dependent endocytosis

with pharmacologic agents and overexpressed dominant-negative proteins. Valuable results have indicated

various CoV entry pathways; caveolar endocytosis for human 229E-CoV [69], clathrin-mediated

endocytosis for SARS-CoV [42], cholesterol-raft entry for MHV [70], clathrin and caveolar–

independent endocytosis for feline CoV [71]. However, complete blockade of CoV entry is rarely

achieved by suppressing any single endocytic pathway, suggesting that multiple cell entry routes can

be utilized by the CoVs. That CoVs can resourcefully invade cells in alternate routes is exemplified by

successful SARS-CoV entry when both clathrin and caveolar endocytosis is blocked [72]. Similarly,

SARS-CoV entry, an ACE2 receptor-mediated process that typically traffics viruses into acidic

organelles [69], is entirely distinct when mediated by antibodies and Fc-gamma-receptor II. In this

so-called antibody-dependent enhanced entry, infection is entirely unaffected by the endosome

-neutralizing agents and protease inhibitors that arrest the ACE2-mediated entry process [73]. In

summary, different CoV endocytic entry routes can culminate in successful infections.

In the endosomes, CoVs are bathed in an acidic and often proteolytic environment. The S proteins

from the majority of CoV strains clearly require acid exposures to acquire membrane fusion

competence. However, it is unclear whether protons are the endosomal factor(s) triggering fusion per

se, as acidification itself does not enhance S-mediated fusions [29]. Acid pH-activated endosomal

proteases may be the more likely triggers [74-76]. Hence an interesting question arises as to whether

entry depends on acidic endosomal conditions solely because the fusion-activating protease(s) are

themselves acid pH-dependent. This is a question considered in studies of filovirus entry [77] and the

CoVs present an additional study model here because the effects of protonation on CoV S fusion

structures can be dynamically evaluated by magnetic resonance [78]. The CoV model system appears

poised to contribute to a general understanding of proton-triggers of virus entry.

5. Virus Entry, S Proteolysis and Membrane Fusion

The initial and most obvious S proteolysis takes place in virus-producing cells, at the “CS-1”

position (Figure 1), to generate secreted viruses with peripheral (S1) and transmembrane (S2)

fragments (event “1” in Figure 3). As stated in the previous section on S morphogenesis, these S1-S2

scissions are not observed in all CoVs. For those CoVs with uncleaved S proteins, the relevant

scissions take place during virus entry (events “3” and “4” in Figure 3). Simmons et al. [76] and

Matsuyama et al. [79] made the initial seminal discovery that cathepsin L inhibitors block SARS S

mediated virus entry and that trypsin proteolysis bypasses this inhibition, making it clear that cell entry

via SARS S proteins depends on a proteolytic cleavage by cathepsin L. These were enlightening

results that arose nearly concurrently with the discovery that Ebolavirus entry requires cathepsin

cleavage of its fusion glycoproteins [80], thus establishing that endosomal proteases can promote the

entry of different enveloped viruses. Subsequent questions regarding the S cleavages were addressed

biochemically, and it was found that SARS S ectodomains were cleaved by cathepsin L at the S1-S2

position (CS1 in Figure 1) [81]. This made it appear that, depending on the CoV strain, S cleavage at

CS1 either takes place early in the infection cycle, by furin-like serine proteases in the trans-Golgi

Page 12: 2012 Ready, Set, Fuse! The Coronavirus Spike Protein and Acquisition of Fusion Competence

Viruses 2012, 4 568

network of virus-producing cells, or later in the cycle, by cathepsin L cysteine proteases in the

endosomes of virus-target cells, with similar ultimate outcomes. Supporting this view, precleaving

the SARS S by inserting a consensus furin cleavage site 1 freed the virus from its dependence on

endosomal acidification [82]. More recent findings, however, have added complexity to these S

proteolytic cleavages. An additional mutibasic peptide motif (at a position we designated in Figure 1 as

“CS2”) was recognized in infectious bronchitis CoV (IBV) S, and shown via mutagenesis to be

essential for IBV cell entry [83]. The homologous position in SARS S was also found to be highly

relevant, as its mutation into a consensus furin cleavage motif allowed for SARS S–mediated entry

without the need for exogenous trypsin proteases [84]. This second cleavage site (CS2) has since been

biochemically confirmed via identification of the shorter C-terminal S fragment, both with SARS S

proteins [85] and with MHV S proteins [57]. Although research on S proteolytic cascades is ongoing,

the present findings are consistent with a model in which S proteins are cleaved multiply, likely first at

the CS1 (S1-S2) position, and then at CS2 positions adjacent to the FP, to generate fusion intermediate

structures of the type depicted in Figure 2. The CS2 proteolytic removal of sequences that are

N-terminal to the FP may permit the rapid completion of the S refolding events that generate

membrane fusion and the 6-HB configuration.

There is also the question of precisely when the S proteins are available for proteolysis. The weight

of current evidence argues that the target cell, not the free extracellular space, is the site of S

proteolysis. For example, cell-free SARS-CoV virions were rendered non-infectious by trypsin

pre-treatment of free virions [72]. However, the same trypsin exposures enhanced infectivities of

cell-bound viruses [76,79] (event “3” in Figure 3). These findings suggest that it is the cell

receptor-associated S proteins that are in unique protease-responsive structural forms. While these

experiments used a protease that is not encountered by incoming respiratory CoVs such as SARS-CoV,

the findings still do imply that S conformations change significantly upon receptor interaction or

endocytosis, and in turn illuminate the importance of S conformation and subcellular location to the

effects of S proteolysis.

That exogenously added trypsin could increase the infectivities of cell receptor-associated

SARS-CoV suggests that cell-bound ectoprotease(s) might be the natural cofactors for CoV entry.

Indeed, CoV S can be primed for fusion by recently described cell-surface proteases known as Type II

Transmembrane Serine Proteases (TTSPs) [85-89]. The TTSPs may reside in plasma membrane

microdomains, although this possibility is not yet validated, and they may be available to incoming

viruses at the receptor-interaction stage, to cleave at the CS1 and/or CS2 positions.

The prototype TTSP, enteropeptidase, was discovered nearly a century ago, and the most recent

decades of research have revealed at least 20 additional family members [90,91]. Family member

variations are in the domains between C-terminal protease and N-terminal transmembrane anchors.

The roles of these domains in viral glycoprotein cleavage, if any, are unknown, but they may dictate

substrate specificities by influencing protease subcellular localization or association with other

membrane proteins [90]. Only a subset of TTSPs have been evaluated as mediators of virus entry.

Principal amongst these evaluated TTSPs is transmembrane protease, serine 2 (TMPRSS2). First

described in 1997 [92], a role for TMPRSS2 in normal physiology remains elusive, largely because

TMPRSS2 knockout mice develop normally [93]. But TMPRSS2 may have a fundamental

organism-wide role, as it is expressed in tissues such as the heart, liver, prostate, intestines, and most

Page 13: 2012 Ready, Set, Fuse! The Coronavirus Spike Protein and Acquisition of Fusion Competence

Viruses 2012, 4 569

notably, the lung [92]. Abundance in the lung has prompted screening of respiratory viruses for

TMPRSS2 activation of their glycoproteins. TMPRSS2 can cleave the hemagglutinins (HAs) of

influenza viruses, leading to their increased replication, including the highly pathogenic 1918

strain [94,95]. TMPRSS2 also cleaves the membrane glycoprotein of SARS-CoV. Matsuyama et al.

discovered that increased levels of TMPRSS2 resulted in more cell death and increased SARS-CoV

viral replication in vitro [86]. Shulla et al. confirmed and expanded this finding by noting that

TMPRSS2 specifically increases SARS-CoV entry using pseudotyped reporter assays [88]. TMPRSS2

cleavage may also contribute to immune evasion. By cleaving S from infected-cell surfaces,

TMPRSS2 releases S fragments that bind and incapacitate neutralizing antibodies [87]. Hence the

in vivo infection process may be heavily influenced by TMPRSS2 and related family members, both at

virus entry and release, influencing pathogenesis and immune response.

Another TTSP, Human Airway Trypsin-like Protease (HAT or TMPRSS11d), has brought out

enlightening details concerning member-specific proteolytic properties. In the context of influenza

HA cleavage, HAT has a broader cleavage capacity than TMPRSS2, proteolyzing HA both in

virus-producing cells and in progeny viruses bound to target cell receptors [96]. Thus HAT, not

TMPRSS2, is the more relevant protease operating on influenza at the virus entry stage. In the context

of SARS-CoV and S cleavage, HAT again exhibits a broader cleavage capacity than TMPRSS2,

making it so that HAT can cleave and enhance S-mediated virus entry either in virus-producing

cells or on the surface of virus-target cells [89]. However, overexpressed TMPRSS2 bypasses the

requirement for endosomal acidification and therefore cathepsin activation [86,88], but HAT does not

similarly replace cathepsins in SARS-CoV entry [89]. Thus a further dissection of the various TTSP

substrate specificities will be necessary to precisely identify those most relevant to virus infection,

and efforts in this regard are continuing. For example, the first paper to examine TTSPs in the context

of SARS entry found that TMPRSS11a was capable of slightly enhancing SARS S bearing

pseudoparticles [85]. Subsequent findings indicated that, while TMPRSS11a was capable of modestly

increasing SARS entry at low levels of the protease, TMPRSS2 was a more potent activator of

entry [88]. Most recently, various TTSPs including TMPRSS3, TMPRSS4, TMPRSS6, and Hepsin,

have been evaluated, yet none have exceeded TMPRSS2 in augmenting SARS-CoV entry [87,89].

Other candidate TTSPs worth testing in SARS-CoV entry assays are MSPL and TMPRSS13, as they

have been found to cleave certain influenza HAs [97].

While the TTSPs may be the most relevant proteases in natural CoV infections, they are clearly

dispensable in several tissue culture settings. This is because cathepsins, specifically cathepsin L, will

proteolytically activate SARS CoV S proteins following virus endocytosis (event “4” in Figure 3)

Multiple proteases with possibly redundant virus entry functions make it difficult to discern which

proteases are necessary for viral entry. This difficulty is perhaps most recognized by the fact that the

presumed proteolytic activation of another human CoV, NL63, is entirely unclear. NL63 S-mediated

entry was not affected by preventing endosomal acidification or by cathepsin inhibitors [98]. While

NL63 is similar to SARS in that it binds to the same receptor, ACE2, the protease responsible for

NL63 S cleavage remains a mystery.

Page 14: 2012 Ready, Set, Fuse! The Coronavirus Spike Protein and Acquisition of Fusion Competence

Viruses 2012, 4 570

6. Therapeutics and Inhibition of CoV Entry

Currently, vaccination is the best clinical approach to eradicating viruses. Experimental SARS-CoV

vaccine immunogens include inactivated whole viruses, entire S trimers, and RBDs (reviewed in [99]).

All formulations elicit protective antibodies that prevent virus entry, some protecting against

lethal SARS-CoV challenge, with most antibodies operating sterically to block virus-receptor

engagement [100,101]. However, protective antibodies can operate variably and in novel ways, as

shown by a human monoclonal antibody directed against the trypsin cleavage site in S that reduced

viral titers and pathology associated with SARS-CoV infection [102]. Interestingly, despite binding to

a proteolytic substrate site, this antibody did not prevent S cleavage but rather prevented membrane

fusion by an unknown mechanism [102]. The antibody was created in transgenic mice that produce

human immunoglobulins, representing an interesting new production method for passive human

immunization. These new approaches signal subunit vaccination and therapeutic antibodies as attractive

future clinical options, but the methods must be evaluated cautiously, as anti-SARS antibodies can

permit virus entry via Fcγ receptors, bypassing the ACE2-dependent entry route and actually

enhancing disease [73].

Attractive natural-product anti-CoV agents include lectins that bind avidly to the unprocessed,

high-mannose carbohydrates on CoV S proteins. One such lectin is griffithsin, an oligomer with six

high-affinity mannose binding sites. Griffithsin is known to block HIV infection via envelope protein

binding [103]. Griffithsin was found to have analogous anti-SARS activity, preventing SARS-CoV

infection and pathology both in vitro and in a mouse disease model, without incidental cytotoxity [104].

Other potential anti-human CoV lectins include the mannose-binding lectin, which arrests SARS-CoV

entry [105] and the plant lectins Galanthus nivalis agglutinin, Hippeastrum hybrid agglutinin, and

Urtica dioica agglutinin, which, to date, have only been evaluated for anti-MHV activity [106].

Notably, most of these lectins do not affect binding of S to the receptor, but rather seem to inhibit entry

at a later stage [105,106].

Several of the proteases described in this paper are potential antiviral therapeutic targets Indeed,

furin-inhibiting peptides arrest cell-cell fusions mediated by coronaviruses MHV and IBV (infectious

bronchitis virus) [36,83]. A novel druglike cathepsin L inhibitor, identified by high-throughput

screening, prevents Ebola and SARS-CoV cell entry [107]. TTSPs also promise to be antiviral drug

targets. With their proteolytically active sites exposed to the extracellular milieu, hydrophilic inhibitors

that remain excluded from cells may be suitable, thus reducing medicinal chemistry and toxicity

problems. Transient, pharmacologic TMPRSS2 suppression may have no health consequences, given

that the TMPRSS2-null mouse is phenotypically normal. Targeting other cellular proteases, such as

furins or cathepsins, may have unintended or toxic side effects due to their importance in normal

physiology. Finally, targeting host proteins is a sensible adjunct to the design of inhibitors targeted

toward highly-mutable virus protein targets.

In addition to preventing CoV entry, inhibiting TTSPs and other proteases would also prevent the

maturation of other viral glycoproteins, blocking entry of a variety of viruses. An example of a

potentially multi or pan-viral protease inhibitor is the human cathepsin L inihibitor tetrahydroquinoline

oxocarbazate that blocks both SARS and Ebola glycoprotein mediated entry in vitro [107]. While

animal studies have yet to be published on that cathepsin inhibitor, Bahgat et al. recently reported that

Page 15: 2012 Ready, Set, Fuse! The Coronavirus Spike Protein and Acquisition of Fusion Competence

Viruses 2012, 4 571

a cocktail of serine protease inhibitors reduced influenza infection and subsequent disease in

mice [108]. Additionally, more historic reports used single agent serine protease inhibitors such as

camostat, leupeptin and aprotinin to decrease influenza infection both in mice and in humans [109,110].

Zihnov et al. reported that aerosol delivery of aprotinin shortened influenza symptomatology in

humans [110]. It is possible that the antiviral aprotinin target was TMPRSS2, given that TMPRSS2 is

abundant in lungs and that aprotinin cannot penetrate the plasma membrane. In support of this

possibility, TMPRSS2 knockdown with peptide-conjugated morpholino oligomers, which are stable

antisense oligonucleotides, rendered TMPRSS2 positive human airway cells resistant to influenza

virus replication [111].

In the end, inhibiting human CoVs may require a multi-target approach. Blocking multiple

proteases along the entry pathway might be combined with lectins, and also with peptide-based fusion

inhibitors. Fusion of the viral membrane with the target cell membrane can be arrested by peptides

mimicking the S heptad repeat region 2 (HR2), which bind and inactivate fusion intermediate S

forms [26], as mentioned earlier in this review. Different HR2 peptides have been investigated with

IC50s in the nanomolar to micromolar range [26,112]; these high affinities and specificities are similar

to the HIV HR2 peptide, enfuvirtide, currently used in humans [113]. Further refinement of the HR2

peptides, and development of peptidomimetics, might enhance antiviral activities and make them even

more suitable for anti-CoV treatment [27]. HR2 peptides cover a broad viral target and thus impose

barriers against virus mutational escape routes. Mutations in HR1 that both confer HR2 resistance and

preserve virus viability are rare [114]. While HR2 peptides could be extremely potent entry inhibitors,

our understanding of SARS entry makes multiple therapeutics targeting viral binding and fusion

feasible and most likely of great value. Certainly, a multifaceted approach to SARS entry inhibition

would be the most likely to succeed.

7. Future Directions

The CoV S peplomers directing cell entry are the principal viral components communicating with

diverse extracellular environments. These arrays of extracellular selective pressures fix diversity into

the genetically-plastic CoV S proteins. Novel S structures allow the CoVs to expand into different

ecological niches; zoonotic emergence of the SARS-CoV being the prime example of this emergence

and its consequences. Understanding S variation, structure, and entry mechanisms are therefore central

to combating potential epidemic CoV transmissions.

The natural diversity of the CoV S proteins provides insights into evolution and mechanism of cell

entry. For example, amongst the CoVs, there are central distinctions in RBDs and fusion–activating

triggers. These differences can dramatically influence pathogenicity by redirecting cell tropism, viral entry

routes and rates. The CoVs are facile models to correlate variations in cell entry with disease potential.

The CoVs are also excellent tools to reveal new host cell components facilitating entry. Of

relevance to disease are the possible species and organ-specific variations in CoV receptors and

S-activating proteases. One of the next key steps in CoV research will be to move toward in vivo

dissection of entry mediators and relate disease manifestations to their presence and abundance in

distinct tissues and species.

Page 16: 2012 Ready, Set, Fuse! The Coronavirus Spike Protein and Acquisition of Fusion Competence

Viruses 2012, 4 572

Increasingly sophisticated in vivo CoV models will also be useful in evaluating new antiviral

strategies. Notably, the infection models and the prototype antiviral agents promise to deliver therapies

that will both guard against potentially serious zoonotic CoVs and reveal more broad–spectrum

antivirals. For example, TTSP inhibitors may be suitable therapeutics for a variety of respiratory

viruses that rely on proteolysis for cell entry. CoV research will continue to reveal features of virus

operating mechanisms and clinically applicable antiviral strategies for the foreseeable future.

Conflict of Interest

The authors declare no conflict of interest.

References

1. Opstelten, D.J.; de Groote, P.; Horzinek, M.C.; Vennema, H.; Rottier, P.J. Disulfide bonds in

folding and transport of mouse hepatitis coronavirus glycoproteins. J. Virol. 1993, 67, 7394–7401.

2. Tooze, J.; Tooze, S.; Warren, G. Replication of coronavirus mhv-a59 in sac- cells: Determination

of the first site of budding of progeny virions. Eur. J. Cell Biol. 1984, 33, 281–293.

3. Neuman, B.W.; Adair, B.D.; Yoshioka, C.; Quispe, J.D.; Milligan, R.A.; Yeager, M.;

Buchmeier, M.J. Ultrastructure of sars-cov, fipv, and mhv revealed by electron cryomicroscopy.

Adv. Exp. Med. Biol. 2006, 581, 181–185.

4. Neuman, B.W.; Adair, B.D.; Yoshioka, C.; Quispe, J.D.; Orca, G.; Kuhn, P.; Milligan, R.A.;

Yeager, M.; Buchmeier, M.J. Supramolecular architecture of severe acute respiratory syndrome

coronavirus revealed by electron cryomicroscopy. J. Virol. 2006, 80, 7918–7928.

5. Beniac, D.R.; Andonov, A.; Grudeski, E.; Booth, T.F. Architecture of the sars coronavirus

prefusion spike. Nat. Struct. Mol. Biol. 2006, 13, 751–752.

6. Wu, K.; Li, W.; Peng, G.; Li, F. Crystal structure of nl63 respiratory coronavirus

receptor-binding domain complexed with its human receptor. Proc. Natl. Acad. Sci. USA 2009,

106, 19970–19974.

7. Li, F.; Li, W.; Farzan, M.; Harrison, S.C. Structure of sars coronavirus spike receptor-binding

domain complexed with receptor. Science 2005, 309, 1864–1868.

8. Peng, G.; Sun, D.; Rajashankar, K.R.; Qian, Z.; Holmes, K.V.; Li, F. Crystal structure of mouse

coronavirus receptor-binding domain complexed with its murine receptor. Proc. Natl. Acad. Sci.

USA 2011, 108, 10696–10701.

9. Wu, K.; Peng, G.; Wilken, M.; Geraghty, R.J.; Li, F. Mechanisms of host receptor adaptation by

sars coronavirus. J. Biol. Chem. 2012, 287, 8904–8911.

10. Xu, Y.; Zhu, J.; Liu, Y.; Lou, Z.; Yuan, F.; Liu, Y.; Cole, D.K.; Ni, L.; Su, N.; Qin, L.; et al.

Characterization of the heptad repeat regions, hr1 and hr2, and design of a fusion core structure

model of the spike protein from severe acute respiratory syndrome (sars) coronavirus.

Biochemistry 2004, 43, 14064–14071.

11. Deng, Y.; Liu, J.; Zheng, Q.; Yong, W.; Lu, M. Structures and polymorphic interactions of two

heptad-repeat regions of the sars virus s2 protein. Structure 2006, 14, 889–899.

Page 17: 2012 Ready, Set, Fuse! The Coronavirus Spike Protein and Acquisition of Fusion Competence

Viruses 2012, 4 573

12. Xu, Y.; Liu, Y.; Lou, Z.; Qin, L.; Li, X.; Bai, Z.; Pang, H.; Tien, P.; Gao, G.F.; Rao, Z. Structural

basis for coronavirus-mediated membrane fusion. Crystal structure of mouse hepatitis virus spike

protein fusion core. J. Biol. Chem. 2004, 279, 30514–30522.

13. Zheng, Q.; Deng, Y.; Liu, J.; van der Hoek, L.; Berkhout, B.; Lu, M. Core structure of s2 from

the human coronavirus nl63 spike glycoprotein. Biochemistry 2006, 45, 15205–15215.

14. Wu, K.; Chen, L.; Peng, G.; Zhou, W.; Pennell, C.A.; Mansky, L.M.; Geraghty, R.J.; Li, F. A

virus-binding hot spot on human angiotensin-converting enzyme 2 is critical for binding of two

different coronaviruses. J. Virol. 2011, 85, 5331–5337.

15. Duquerroy, S.; Vigouroux, A.; Rottier, P.J.; Rey, F.A.; Bosch, B.J. Central ions and lateral

asparagine/glutamine zippers stabilize the post-fusion hairpin conformation of the sars

coronavirus spike glycoprotein. Virology 2005, 335, 276–285.

16. Lamb, R.A.; Jardetzky, T.S. Structural basis of viral invasion: Lessons from paramyxovirus f.

Curr. Opin. Struct. Biol. 2007, 17, 427–436.

17. Luo, Z.; Weiss, S.R. Roles in cell-to-cell fusion of two conserved hydrophobic regions in the

murine coronavirus spike protein. Virology 1998, 244, 483–494.

18. Madu, I.G.; Roth, S.L.; Belouzard, S.; Whittaker, G.R. Characterization of a highly conserved

domain within the severe acute respiratory syndrome coronavirus spike protein s2 domain with

characteristics of a viral fusion peptide. J. Virol. 2009, 83, 7411–7421.

19. Broer, R.; Boson, B.; Spaan, W.; Cosset, F.L.; Corver, J. Important role for the transmembrane

domain of severe acute respiratory syndrome coronavirus spike protein during entry. J. Virol.

2006, 80, 1302–1310.

20. Lu, Y.; Neo, T.L.; Liu, D.X.; Tam, J.P. Importance of sars-cov spike protein trp-rich region in

viral infectivity. Biochem. Biophys. Res. Commun. 2008, 371, 356–360.

21. Bos, E.C.; Heijnen, L.; Luytjes, W.; Spaan, W.J. Mutational analysis of the murine coronavirus

spike protein: Effect on cell-to-cell fusion. Virology 1995, 214, 453–463.

22. Petit, C.M.; Chouljenko, V.N.; Iyer, A.; Colgrove, R.; Farzan, M.; Knipe, D.M.; Kousoulas, K.G.

Palmitoylation of the cysteine-rich endodomain of the sars-coronavirus spike glycoprotein is

important for spike-mediated cell fusion. Virology 2007, 360, 264–274.

23. Shulla, A.; Gallagher, T. Role of spike protein endodomains in regulating coronavirus entry.

J. Biol. Chem. 2009, 284, 32725–32734.

24. Bosch, B.J.; van der Zee, R.; de Haan, C.A.; Rottier, P.J. The coronavirus spike protein is a class

i virus fusion protein: Structural and functional characterization of the fusion core complex.

J. Virol. 2003, 77, 8801–8811.

25. Tripet, B.; Howard, M.W.; Jobling, M.; Holmes, R.K.; Holmes, K.V.; Hodges, R.S. Structural

characterization of the sars-coronavirus spike s fusion protein core. J. Biol. Chem. 2004, 279,

20836–20849.

26. Bosch, B.J.; Martina, B.E.; Van Der Zee, R.; Lepault, J.; Haijema, B.J.; Versluis, C.; Heck, A.J.;

De Groot, R.; Osterhaus, A.D.; Rottier, P.J. Severe acute respiratory syndrome coronavirus

(sars-cov) infection inhibition using spike protein heptad repeat-derived peptides. Proc. Natil.

Acad. Sci. USA 2004, 101, 8455–8460.

Page 18: 2012 Ready, Set, Fuse! The Coronavirus Spike Protein and Acquisition of Fusion Competence

Viruses 2012, 4 574

27. Liu, I.J.; Kao, C.L.; Hsieh, S.C.; Wey, M.T.; Kan, L.S.; Wang, W.K. Identification of a minimal

peptide derived from heptad repeat (hr) 2 of spike protein of sars-cov and combination of

hr1-derived peptides as fusion inhibitors. Antiviral Res. 2009, 81, 82–87.

28. Zheng, B.J.; Guan, Y.; Hez, M.L.; Sun, H.; Du, L.; Zheng, Y.; Wong, K.L.; Chen, H.; Chen, Y.;

Lu, L.; et al. Synthetic peptides outside the spike protein heptad repeat regions as potent

inhibitors of sars-associated coronavirus. Antiviral Ther. 2005, 10, 393–403.

29. Sturman, L.S.; Ricard, C.S.; Holmes, K.V. Conformational change of the coronavirus peplomer

glycoprotein at ph 8.0 and 37 degrees c correlates with virus aggregation and virus-induced cell

fusion. J. Virol. 1990, 64, 3042–3050.

30. Gallagher, T.M. A role for naturally occurring variation of the murine coronavirus spike protein

in stabilizing association with the cellular receptor. J. Virol. 1997, 71, 3129–3137.

31. Ruch, T.R.; Machamer, C.E. The hydrophobic domain of infectious bronchitis virus e protein

alters the host secretory pathway and is important for release of infectious virus. J. Virol. 2011,

85, 675–685.

32. Nieto-Torres, J.L.; Dediego, M.L.; Alvarez, E.; Jimenez-Guardeno, J.M.; Regla-Nava, J.A.;

Llorente, M.; Kremer, L.; Shuo, S.; Enjuanes, L. Subcellular location and topology of severe

acute respiratory syndrome coronavirus envelope protein. Virology 2011, 415, 69–82.

33. Wilson, L.; McKinlay, C.; Gage, P.; Ewart, G. Sars coronavirus e protein forms cation-selective

ion channels. Virology 2004, 330, 322–331.

34. Ortego, J.; Ceriani, J.E.; Patino, C.; Plana, J.; Enjuanes, L. Absence of e protein arrests

transmissible gastroenteritis coronavirus maturation in the secretory pathway. Virology 2007,

368, 296–308.

35. Feliciangeli, S.F.; Thomas, L.; Scott, G.K.; Subbian, E.; Hung, C.H.; Molloy, S.S.; Jean, F.;

Shinde, U.; Thomas, G. Identification of a ph sensor in the furin propeptide that regulates

enzyme activation. J. Biol. Chem. 2006, 281, 16108–16116.

36. de Haan, C.A.; Stadler, K.; Godeke, G.J.; Bosch, B.J.; Rottier, P.J. Cleavage inhibition of the

murine coronavirus spike protein by a furin-like enzyme affects cell-cell but not virus-cell fusion.

J. Virol. 2004, 78, 6048–6054.

37. Gombold, J.L.; Hingley, S.T.; Weiss, S.R. Fusion-defective mutants of mouse hepatitis virus a59

contain a mutation in the spike protein cleavage signal. J. Virol. 1993, 67, 4504–4512.

38. Yamada, Y.K.; Takimoto, K.; Yabe, M.; Taguchi, F. Acquired fusion activity of a murine

coronavirus mhv-2 variant with mutations in the proteolytic cleavage site and the signal sequence

of the s protein. Virology 1997, 227, 215–219.

39. Frana, M.F.; Behnke, J.N.; Sturman, L.S.; Holmes, K.V. Proteolytic cleavage of the e2

glycoprotein of murine coronavirus: Host-dependent differences in proteolytic cleavage and cell

fusion. J. Virol. 1985, 56, 912–920.

40. Follis, K.E.; York, J.; Nunberg, J.H. Furin cleavage of the sars coronavirus spike glycoprotein

enhances cell-cell fusion but does not affect virion entry. Virology 2006, 350, 358–369.

41. de Haan, C.A.; Li, Z.; te Lintelo, E.; Bosch, B.J.; Haijema, B.J.; Rottier, P.J. Murine coronavirus

with an extended host range uses heparan sulfate as an entry receptor. J. Virol. 2005, 79,

14451–14456.

Page 19: 2012 Ready, Set, Fuse! The Coronavirus Spike Protein and Acquisition of Fusion Competence

Viruses 2012, 4 575

42. Inoue, Y.; Tanaka, N.; Tanaka, Y.; Inoue, S.; Morita, K.; Zhuang, M.; Hattori, T.; Sugamura, K.

Clathrin-dependent entry of severe acute respiratory syndrome coronavirus into target cells

expressing ace2 with the cytoplasmic tail deleted. J. Virol. 2007, 81, 8722–8729.

43. Thorp, E.B.; Gallagher, T.M. Requirements for ceacams and cholesterol during murine

coronavirus cell entry. J. Virol. 2004, 78, 2682–2692.

44. Watanabe, R.; Matsuyama, S.; Taguchi, F. Receptor-independent infection of murine

coronavirus: Analysis by spinoculation. J. Virol. 2006, 80, 4901–4908.

45. Miura, H.S.; Nakagaki, K.; Taguchi, F. N-terminal domain of the murine coronavirus receptor

ceacam1 is responsible for fusogenic activation and conformational changes of the spike protein.

J. Virol. 2004, 78, 216–223.

46. Matsuyama, S.; Taguchi, F. Receptor-induced conformational changes of murine coronavirus

spike protein. J. Virol. 2002, 76, 11819–11826.

47. Li, W.; Zhang, C.; Sui, J.; Kuhn, J.H.; Moore, M.J.; Luo, S.; Wong, S.K.; Huang, I.C.; Xu, K.;

Vasilieva, N.; et al. Receptor and viral determinants of sars-coronavirus adaptation to human

ace2. EMBO J. 2005, 24, 1634–1643.

48. Becker, M.M.; Graham, R.L.; Donaldson, E.F.; Rockx, B.; Sims, A.C.; Sheahan, T.;

Pickles, R.J.; Corti, D.; Johnston, R.E.; Baric, R.S.; et al. Synthetic recombinant bat sars-like

coronavirus is infectious in cultured cells and in mice. Proc. Natl. Acad. Sci. USA 2008, 105,

19944–19949.

49. Gallagher, T.M.; Parker, S.E.; Buchmeier, M.J. Neutralization-resistant variants of a neurotropic

coronavirus are generated by deletions within the amino-terminal half of the spike glycoprotein.

J. Virol. 1990, 64, 731–741.

50. Rowe, C.L.; Fleming, J.O.; Nathan, M.J.; Sgro, J.Y.; Palmenberg, A.C.; Baker, S.C. Generation

of coronavirus spike deletion variants by high-frequency recombination at regions of predicted

rna secondary structure. J. Virol. 1997, 71, 6183–6190.

51. Phillips, J.M.; Weiss, S.R. Pathogenesis of neurotropic murine coronavirus is multifactorial.

Trends Pharmacol. Sci. 2011, 32, 2–7.

52. Enjuanes, L.; Sanchez, C.; Gebauer, F.; Mendez, A.; Dopazo, J.; Ballesteros, M.L. Evolution and

tropism of transmissible gastroenteritis coronavirus. Adv. Exp. Med. Biol. 1993, 342, 35–42.

53. Schultze, B.; Krempl, C.; Ballesteros, M.L.; Shaw, L.; Schauer, R.; Enjuanes, L.; Herrler, G.

Transmissible gastroenteritis coronavirus, but not the related porcine respiratory coronavirus, has

a sialic acid (n-glycolylneuraminic acid) binding activity. J. Virol. 1996, 70, 5634–5637.

54. Sanchez, C.M.; Gebauer, F.; Sune, C.; Mendez, A.; Dopazo, J.; Enjuanes, L. Genetic evolution

and tropism of transmissible gastroenteritis coronaviruses. Virology 1992, 190, 92–105.

55. Ballesteros, M.L.; Sanchez, C.M.; Enjuanes, L. Two amino acid changes at the n-terminus of

transmissible gastroenteritis coronavirus spike protein result in the loss of enteric tropism.

Virology 1997, 227, 378–388.

56. Zelus, B.D.; Schickli, J.H.; Blau, D.M.; Weiss, S.R.; Holmes, K.V. Conformational changes in

the spike glycoprotein of murine coronavirus are induced at 37 degrees c either by soluble

murine ceacam1 receptors or by ph 8. J. Virol. 2003, 77, 830–840.

57. Matsuyama, S.; Taguchi, F. Two-step conformational changes in a coronavirus envelope

glycoprotein mediated by receptor binding and proteolysis. J. Virol. 2009, 83, 11133–11141.

Page 20: 2012 Ready, Set, Fuse! The Coronavirus Spike Protein and Acquisition of Fusion Competence

Viruses 2012, 4 576

58. Grosse, B.; Siddell, S.G. Single amino acid changes in the s2 subunit of the mhv surface

glycoprotein confer resistance to neutralization by s1 subunit-specific monoclonal antibody.

Virology 1994, 202, 814–824.

59. Ontiveros, E.; Kim, T.S.; Gallagher, T.M.; Perlman, S. Enhanced virulence mediated by the

murine coronavirus, mouse hepatitis virus strain jhm, is associated with a glycine at residue 310

of the spike glycoprotein. J. Virol. 2003, 77, 10260–10269.

60. de Haan, C.A.; Te Lintelo, E.; Li, Z.; Raaben, M.; Wurdinger, T.; Bosch, B.J.; Rottier, P.J.

Cooperative involvement of the s1 and s2 subunits of the murine coronavirus spike protein in

receptor binding and extended host range. J. Virol. 2006, 80, 10909–10918.

61. Matsuyama, S.; Taguchi, F. Communication between s1n330 and a region in s2 of murine

coronavirus spike protein is important for virus entry into cells expressing ceacam1b receptor.

Virology 2002, 295, 160–171.

62. Krueger, D.K.; Kelly, S.M.; Lewicki, D.N.; Ruffolo, R.; Gallagher, T.M. Variations in disparate

regions of the murine coronavirus spike protein impact the initiation of membrane fusion.

J. Virol. 2001, 75, 2792–2802.

63. Sieczkarski, S.B.; Whittaker, G.R. Viral entry. Curr. Top. Microbiol. Immunol. 2005, 285, 1–23.

64. Marsh, M.; Bron, R. Sfv infection in cho cells: Cell-type specific restrictions to productive virus

entry at the cell surface. J. Cell Sci. 1997, 110 (Pt 1), 95–103.

65. Gallagher, T.M.; Escarmis, C.; Buchmeier, M.J. Alteration of the ph dependence of coronavirus-

induced cell fusion: Effect of mutations in the spike glycoprotein. J. Virol. 1991, 65, 1916–1928.

66. Roos, D.S.; Duchala, C.S.; Stephensen, C.B.; Holmes, K.V.; Choppin, P.W. Control of

virus-induced cell fusion by host cell lipid composition. Virology 1990, 175, 345–357.

67. Miyauchi, K.; Kim, Y.; Latinovic, O.; Morozov, V.; Melikyan, G.B. Hiv enters cells via

endocytosis and dynamin-dependent fusion with endosomes. Cell 2009, 137, 433–444.

68. Sieczkarski, S.B.; Whittaker, G.R. Dissecting virus entry via endocytosis. J. Gen. Virol. 2002,

83, 1535–1545.

69. Nomura, R.; Kiyota, A.; Suzaki, E.; Kataoka, K.; Ohe, Y.; Miyamoto, K.; Senda, T.; Fujimoto, T.

Human coronavirus 229e binds to cd13 in rafts and enters the cell through caveolae. J. Virol.

2004, 78, 8701–8708.

70. Choi, K.S.; Aizaki, H.; Lai, M.M. Murine coronavirus requires lipid rafts for virus entry and cell-

cell fusion but not for virus release. J. Virol. 2005, 79, 9862–9871.

71. Van Hamme, E.; Dewerchin, H.L.; Cornelissen, E.; Verhasselt, B.; Nauwynck, H.J. Clathrin- and

caveolae-independent entry of feline infectious peritonitis virus in monocytes depends on

dynamin. J. Gen. Virol. 2008, 89, 2147–2156.

72. Wang, H.; Yang, P.; Liu, K.; Guo, F.; Zhang, Y.; Zhang, G.; Jiang, C. Sars coronavirus entry into

host cells through a novel clathrin- and caveolae-independent endocytic pathway. Cell Res. 2008,

18, 290–301.

73. Jaume, M.; Yip, M.S.; Cheung, C.Y.; Leung, H.L.; Li, P.H.; Kien, F.; Dutry, I.; Callendret, B.;

Escriou, N.; Altmeyer, R.; et al. Anti-severe acute respiratory syndrome coronavirus spike

antibodies trigger infection of human immune cells via a ph- and cysteine protease-independent

fcgammar pathway. J. Virol. 2011, 85, 10582–10597.

Page 21: 2012 Ready, Set, Fuse! The Coronavirus Spike Protein and Acquisition of Fusion Competence

Viruses 2012, 4 577

74. Simmons, G.; Reeves, J.D.; Rennekamp, A.J.; Amberg, S.M.; Piefer, A.J.; Bates, P.

Characterization of severe acute respiratory syndrome-associated coronavirus (sars-cov) spike

glycoprotein-mediated viral entry. Proc. Natil. Acad. Sci. USA 2004, 101, 4240–4245.

75. Qiu, Z.; Hingley, S.T.; Simmons, G.; Yu, C.; Das Sarma, J.; Bates, P.; Weiss, S.R. Endosomal

proteolysis by cathepsins is necessary for murine coronavirus mouse hepatitis virus type 2

spike-mediated entry. J. Virol. 2006, 80, 5768–5776.

76. Simmons, G.; Gosalia, D.N.; Rennekamp, A.J.; Reeves, J.D.; Diamond, S.L.; Bates, P. Inhibitors

of cathepsin l prevent severe acute respiratory syndrome coronavirus entry. Proc. Natil. Acad.

Sci. USA 2005, 102, 11876–11881.

77. Gregory, S.M.; Harada, E.; Liang, B.; Delos, S.E.; White, J.M.; Tamm, L.K. Structure and

function of the complete internal fusion loop from ebolavirus glycoprotein 2. Proc. Natil. Acad.

Sci. USA 2011, 108, 11211–11216.

78. McReynolds, S.; Jiang, S.; Guo, Y.; Celigoy, J.; Schar, C.; Rong, L.; Caffrey, M.

Characterization of the prefusion and transition states of severe acute respiratory syndrome

coronavirus s2-hr2. Biochemistry 2008, 47, 6802–6808.

79. Matsuyama, S.; Ujike, M.; Morikawa, S.; Tashiro, M.; Taguchi, F. Protease-mediated

enhancement of severe acute respiratory syndrome coronavirus infection. Proc. Natil. Acad. Sci.

USA 2005, 102, 12543–12547.

80. Chandran, K.; Sullivan, N.J.; Felbor, U.; Whelan, S.P.; Cunningham, J.M. Endosomal proteolysis

of the ebola virus glycoprotein is necessary for infection. Science 2005, 308, 1643–1645.

81. Bosch, B.J.; Bartelink, W.; Rottier, P.J. Cathepsin l functionally cleaves the severe acute

respiratory syndrome coronavirus class i fusion protein upstream of rather than adjacent to the

fusion peptide. J. Virol. 2008, 82, 8887–8890.

82. Watanabe, R.; Matsuyama, S.; Shirato, K.; Maejima, M.; Fukushi, S.; Morikawa, S.; Taguchi, F.

Entry from the cell surface of severe acute respiratory syndrome coronavirus with cleaved s

protein as revealed by pseudotype virus bearing cleaved s protein. J. Virol. 2008, 82, 11985–11991.

83. Yamada, Y.; Liu, D.X. Proteolytic activation of the spike protein at a novel rrrr/s motif is

implicated in furin-dependent entry, syncytium formation, and infectivity of coronavirus

infectious bronchitis virus in cultured cells. J. Virol. 2009, 83, 8744–8758.

84. Belouzard, S.; Chu, V.C.; Whittaker, G.R. Activation of the sars coronavirus spike protein via

sequential proteolytic cleavage at two distinct sites. Proc. Natil. Acad. Sci. USA 2009, 106,

5871–5876.

85. Kam, Y.W.; Okumura, Y.; Kido, H.; Ng, L.F.; Bruzzone, R.; Altmeyer, R. Cleavage of the sars

coronavirus spike glycoprotein by airway proteases enhances virus entry into human bronchial

epithelial cells in vitro. PLoS One 2009, 4, e7870.

86. Matsuyama, S.; Nagata, N.; Shirato, K.; Kawase, M.; Takeda, M.; Taguchi, F. Efficient

activation of the severe acute respiratory syndrome coronavirus spike protein by the

transmembrane protease tmprss2. J. Virol. 2010, 84, 12658–12664.

87. Glowacka, I.; Bertram, S.; Muller, M.A.; Allen, P.; Soilleux, E.; Pfefferle, S.; Steffen, I.;

Tsegaye, T.S.; He, Y.; Gnirss, K.; et al. Evidence that tmprss2 activates the severe acute

respiratory syndrome coronavirus spike protein for membrane fusion and reduces viral control by

the humoral immune response. J. Virol. 2011, 85, 4122–4134.

Page 22: 2012 Ready, Set, Fuse! The Coronavirus Spike Protein and Acquisition of Fusion Competence

Viruses 2012, 4 578

88. Shulla, A.; Heald-Sargent, T.; Subramanya, G.; Zhao, J.; Perlman, S.; Gallagher, T. A

transmembrane serine protease is linked to the severe acute respiratory syndrome coronavirus

receptor and activates virus entry. J. Virol. 2011, 85, 873–882.

89. Bertram, S.; Glowacka, I.; Muller, M.A.; Lavender, H.; Gnirss, K.; Nehlmeier, I.; Niemeyer, D.;

He, Y.; Simmons, G.; Drosten, C.; Soilleux, E.J.; Jahn, O.; Steffen, I.; Pohlmann, S. Cleavage

and activation of the severe acute respiratory syndrome coronavirus spike protein by human

airway trypsin-like protease. J. Virol. 2011, 85, 13363–13372.

90. Szabo, R.; Bugge, T.H. Membrane-anchored serine proteases in vertebrate cell and

developmental biology. Annu. Rev. Cell Dev. Biol. 2011, 27, 213–235.

91. Hooper, J.D.; Clements, J.A.; Quigley, J.P.; Antalis, T.M. Type ii transmembrane serine

proteases. Insights into an emerging class of cell surface proteolytic enzymes. J. Biol. Chem.

2001, 276, 857–860.

92. Paoloni-Giacobino, A.; Chen, H.; Peitsch, M.C.; Rossier, C.; Antonarakis, S.E. Cloning of the

tmprss2 gene, which encodes a novel serine protease with transmembrane, ldlra, and srcr

domains and maps to 21q22.3. Genomics 1997, 44, 309–320.

93. Kim, T.S.; Heinlein, C.; Hackman, R.C.; Nelson, P.S. Phenotypic analysis of mice lacking the

tmprss2-encoded protease. Mol. Cell. Biol. 2006, 26, 965–975.

94. Bottcher, E.; Matrosovich, T.; Beyerle, M.; Klenk, H.D.; Garten, W.; Matrosovich, M.

Proteolytic activation of influenza viruses by serine proteases tmprss2 and hat from human

airway epithelium. J. Virol. 2006, 80, 9896–9898.

95. Chaipan, C.; Kobasa, D.; Bertram, S.; Glowacka, I.; Steffen, I.; Tsegaye, T.S.; Takeda, M.;

Bugge, T.H.; Kim, S.; Park, Y.; Marzi, A.; Pohlmann, S. Proteolytic activation of the 1918

influenza virus hemagglutinin. J. Virol. 2009, 83, 3200–3211.

96. Bottcher-Friebertshauser, E.; Freuer, C.; Sielaff, F.; Schmidt, S.; Eickmann, M.; Uhlendorff, J.;

Steinmetzer, T.; Klenk, H.D.; Garten, W. Cleavage of influenza virus hemagglutinin by airway

proteases tmprss2 and hat differs in subcellular localization and susceptibility to protease

inhibitors. J. Virol. 2010, 84, 5605–5614.

97. Okumura, Y.; Takahashi, E.; Yano, M.; Ohuchi, M.; Daidoji, T.; Nakaya, T.; Bottcher, E.;

Garten, W.; Klenk, H.D.; Kido, H. Novel type ii transmembrane serine proteases, mspl and

tmprss13, proteolytically activate membrane fusion activity of the hemagglutinin of highly

pathogenic avian influenza viruses and induce their multicycle replication. J. Virol. 2010, 84,

5089–5096.

98. Huang, I.C.; Bosch, B.J.; Li, F.; Li, W.; Lee, K.H.; Ghiran, S.; Vasilieva, N.; Dermody, T.S.;

Harrison, S.C.; Dormitzer, P.R.; et al. Sars coronavirus, but not human coronavirus nl63, utilizes

cathepsin l to infect ace2-expressing cells. J. Biol. Chem. 2006, 281, 3198–3203.

99. Du, L.; He, Y.; Zhou, Y.; Liu, S.; Zheng, B.J.; Jiang, S. The spike protein of sars-cov--a target

for vaccine and therapeutic development. Nat. Rev. Microbiol. 2009, 7, 226–236.

100. van den Brink, E.N.; Ter Meulen, J.; Cox, F.; Jongeneelen, M.A.; Thijsse, A.; Throsby, M.;

Marissen, W.E.; Rood, P.M.; Bakker, A.B.; Gelderblom, H.R.; et al. Molecular and biological

characterization of human monoclonal antibodies binding to the spike and nucleocapsid proteins

of severe acute respiratory syndrome coronavirus. J. Virol. 2005, 79, 1635–1644.

Page 23: 2012 Ready, Set, Fuse! The Coronavirus Spike Protein and Acquisition of Fusion Competence

Viruses 2012, 4 579

101. Chen, Z.; Zhang, L.; Qin, C.; Ba, L.; Yi, C.E.; Zhang, F.; Wei, Q.; He, T.; Yu, W.; Yu, J.; et al.

Recombinant modified vaccinia virus ankara expressing the spike glycoprotein of severe acute

respiratory syndrome coronavirus induces protective neutralizing antibodies primarily targeting

the receptor binding region. J. Virol. 2005, 79, 2678–2688.

102. Miyoshi-Akiyama, T.; Ishida, I.; Fukushi, M.; Yamaguchi, K.; Matsuoka, Y.; Ishihara, T.;

Tsukahara, M.; Hatakeyama, S.; Itoh, N.; Morisawa, A.; et al. Fully human monoclonal antibody

directed to proteolytic cleavage site in severe acute respiratory syndrome (sars) coronavirus s

protein neutralizes the virus in a rhesus macaque sars model. J. Infect. Dis. 2011, 203, 1574–1581.

103. Mori, T.; O'Keefe, B.R.; Sowder, R.C., 2nd; Bringans, S.; Gardella, R.; Berg, S.; Cochran, P.;

Turpin, J.A.; Buckheit, R.W., Jr.; McMahon, J.B.; Boyd, M.R. Isolation and characterization of

griffithsin, a novel hiv-inactivating protein, from the red alga griffithsia sp. J. Biol. Chem. 2005,

280, 9345–9353.

104. O'Keefe, B.R.; Giomarelli, B.; Barnard, D.L.; Shenoy, S.R.; Chan, P.K.; McMahon, J.B.;

Palmer, K.E.; Barnett, B.W.; Meyerholz, D.K.; Wohlford-Lenane, C.L.; et al. Broad-spectrum in

vitro activity and in vivo efficacy of the antiviral protein griffithsin against emerging viruses of

the family coronaviridae. J. Virol. 2010, 84, 2511–2521.

105. Zhou, Y.; Lu, K.; Pfefferle, S.; Bertram, S.; Glowacka, I.; Drosten, C.; Pohlmann, S.;

Simmons, G. A single asparagine-linked glycosylation site of the severe acute respiratory

syndrome coronavirus spike glycoprotein facilitates inhibition by mannose-binding lectin

through multiple mechanisms. J. Virol. 2010, 84, 8753–8764.

106. van der Meer, F.J.; de Haan, C.A.; Schuurman, N.M.; Haijema, B.J.; Peumans, W.J.;

van Damme, E.J.; Delputte, P.L.; Balzarini, J.; Egberink, H.F. Antiviral activity of carbohydrate-

binding agents against nidovirales in cell culture. Antiviral Res. 2007, 76, 21–29.

107. Shah, P.P.; Wang, T.; Kaletsky, R.L.; Myers, M.C.; Purvis, J.E.; Jing, H.; Huryn, D.M.;

Greenbaum, D.C.; Smith, A.B., 3rd; Bates, P.; et al. A small-molecule oxocarbazate inhibitor of

human cathepsin l blocks severe acute respiratory syndrome and ebola pseudotype virus infection

into human embryonic kidney 293t cells. Mol. Pharmacol. 2010, 78, 319–324.

108. Bahgat, M.M.; Blazejewska, P.; Schughart, K. Inhibition of lung serine proteases in mice: A

potentially new approach to control influenza infection. Virol. J. 2011, 8, 27.

109. Lee, M.G.; Kim, K.H.; Park, K.Y.; Kim, J.S. Evaluation of anti-influenza effects of camostat in

mice infected with non-adapted human influenza viruses. Arch. Virol. 1996, 141, 1979–1989.

110. Zhirnov, O.P.; Kirzhner, L.S.; Ovcharenko, A.V.; Malyshev, N.A. clinical effectiveness of

aprotinin aerosol in influenza and parainfluenza. Vestnik Rossiiskoi Akademii Meditsinskikh Nauk

/ Rossiiskaia Akademiia Meditsinskikh Nauk 1996, 26–31.

111. Bottcher-Friebertshauser, E.; Stein, D.A.; Klenk, H.D.; Garten, W. Inhibition of influenza virus

infection in human airway cell cultures by an antisense peptide-conjugated morpholino oligomer

targeting the hemagglutinin-activating protease tmprss2. J. Virol. 2011, 85, 1554–1562.

112. Zhu, J.; Xiao, G.; Xu, Y.; Yuan, F.; Zheng, C.; Liu, Y.; Yan, H.; Cole, D.K.; Bell, J.I.; Rao, Z.;

et al. Following the rule: Formation of the 6-helix bundle of the fusion core from severe acute

respiratory syndrome coronavirus spike protein and identification of potent peptide inhibitors.

Biochem. Biophys. Res. Commun. 2004, 319, 283–288.

Page 24: 2012 Ready, Set, Fuse! The Coronavirus Spike Protein and Acquisition of Fusion Competence

Viruses 2012, 4 580

113. Tremblay, C.L.; Kollmann, C.; Giguel, F.; Chou, T.C.; Hirsch, M.S. Strong in vitro synergy

between the fusion inhibitor t-20 and the cxcr4 blocker amd-3100. J. Acquir Immune Defic.

Syndr. 2000, 25, 99–102.

114. Bosch, B.J.; Rossen, J.W.; Bartelink, W.; Zuurveen, S.J.; de Haan, C.A.; Duquerroy, S.;

Boucher, C.A.; Rottier, P.J. Coronavirus escape from heptad repeat 2 (hr2)-derived peptide entry

inhibition as a result of mutations in the hr1 domain of the spike fusion protein. J. Virol. 2008,

82, 2580–2585.

© 2012 by the authors; licensee MDPI, Basel, Switzerland. This article is an open access article

distributed under the terms and conditions of the Creative Commons Attribution license

(http://creativecommons.org/licenses/by/3.0/).